4 Viable Transfer of Microorganisms in the Solar System and Beyond

Page created by Lewis Henderson
 
CONTINUE READING
4 Viable Transfer of Microorganisms in the Solar System and Beyond
4      Viable Transfer of Microorganisms
       in the Solar System and Beyond

Gerda Horneck, Curt Mileikowsky, H. Jay Melosh, John W. Wilson,
Francis A. Cucinotta and Brett Gladman

It is now generally accepted that at the beginning of our solar system a considerable
amount of organic molecules and water were imported to the early Earth as well as to
the other terrestrial planets via asteroids and comets [1-3]. The period of heavy bom-
bardment lasted until approximately 3.8 billion years (Ga) ago. These impactors would
on the one hand have been delivering the volatiles as precursors of life, on the other
hand, if sufficiently large and fast, would have eroded the atmosphere and perhaps
sterilized the Earth and/or Mars, if life existed there [4, 5].
    Impactors of sizes larger than 1 km lead to the ejection of a considerable amount of
soil and rocks that are thrown up at high velocities, some fraction reaching escape
velocity [6]. These ejecta leave the planet and orbit around the Sun, usually for time
scales of a few hundred thousand or several million years until they either impact an-
other celestial body or are expelled out of the solar system [7]. Meteorites of lunar and
some of Martian origin detected within in the last decades are witnesses of these proc-
esses [8, 9]. The question arises whether such rock or soil ejecta could also be the
vehicle for life to leave its planet of origin, or, in other words, whether spreading of
life in the solar system via natural transfer of viable microbes is a feasible process.

4.1 Scenario of Interplanetary Transfer of Life
    Within the Solar System
The supposition that life can be naturally transferred from one planet to another or
even between solar systems goes back to the last century [10] and was formulated as
hypothesis of Panspermia by S. Arrhenius in 1903 [11]. It postulates that microscopic
forms of life, for example spores, can be dispersed in space by the radiation pressure
from the Sun thereby seeding life from one planet to another, or one solar system to
another, respectively. This hypothesis has been subjected to several criticisms with
arguments, such as it cannot be experimentally tested and spores will not survive long-
time exposure to the hostile environment of space, especially vacuum and radiation
(reviewed in [12]). It has also been pointed out that Panspermia only shunts aside the
question of the origin of life to another celestial body (see Chap. 1, Ehrenfreund and
Menten).
   However, a variety of recent discoveries have shed new light on the likelihood of
viable transfer in space such as (i) the detection of meteorites, some of lunar and some
4 Viable Transfer of Microorganisms in the Solar System and Beyond
58      G. Horneck et al.

of Martian origin [9]; (ii) the detection of organics and the still highly debated suppo-
sition of microbial fossils in one of the Martian meteorites [13]; (iii) the probability of
small particles of diameters between 0.5 µm and 1 cm [14], or even boulder-sized
rocks reaching escape velocities by the impact of large comets or asteroids on a planet,
e.g., on Earth [6] or Mars [15, 16]; (iv) the ability of bacterial spores to survive to a
certain extent the shock waves of such a simulated impact [17]; (v) the high UV-
resistance of microorganisms at the low temperatures of deep space, tested at tem-
peratures down to 10 K [18]; (vi) the reported survival of bacterial spores over mil-
lions of years, if enclosed in amber or salt stocks [19, 20], or in space over periods
extending up to 6 years [21]; (vii) the paleogeochemical evidence of a very early ap-
pearance of life on Earth in the form of metabolically advanced microbial prokaryotic
ecosystems leaving not more than approximately 0.4 Ga for the evolution of life from
the simple precursor molecules to the level of a prokaryotic, photoautotrophic cell
[22]; (viii) the biochemical evidence of a common ancestor for all life forms on Earth
[23].
    Viable transfer from one planet to another requires that life, probably of microbial
nature, survives the following three steps: (i) the escape process, i.e. ejection into
space, e.g., caused by a large impact on the parent planet; (ii) the journey through space,
i.e. time scales in space comparable with those experienced by the Martian meteorites
(approximately 1-15 Ma); and (iii) the landing process, i.e. non-destructive deposition
of the biological material on another planet (Fig. 4.1).

            Fig. 4.1 Scenario of an interplanetary transfer of life in the solar system
4 Viable Transfer of Microorganisms in the Solar System and Beyond                59

4.2 Survival of the Escape Process
The most plausible process capable of ejecting microbe-bearing surface material from
a planet or moon into space is the hypervelocity impact of a large object such as an
asteroid or comet, under strong or moderate shock metamorphism of the ejected rock
fragments. The peak shock pressure estimates for the presently studied 15 Martian
meteorites range from about 20 GPa to about 45 GPa and estimates of associated post-
shock temperature range from about 100 °C at 20 GPa to about 600 °C at 45 GPa
[24]. Although these impacts are very energetic, a certain fraction of ejecta is not
heated above 100 °C. These low temperature fragments are ejected from the so-called
spall zone, i.e. the surface layer of the target where the resulting shock is considerably
reduced by superimposition of the reflected shock wave on the direct one [25]. Esti-
mates suggest that within the last 4 Ga, more than 109 fragments of a diameter of ≥2 m
and temperatures ≤100 °C were ejected from Mars of which about 5% arrived on
Earth after a journey in space of ≤8 Ma (Fig. 4.2). The corresponding numbers for a
transfer from Earth to Mars are about 108 fragments ejected from the Earth with about
0.1% arriving on Mars within 8 Ma [26]. During the preceding period of “heavy bom-
bardment” even 10 times higher numbers are estimated. Hence, the 15 Martian mete-
orites, so far detected on Earth, represent probably only an infinitesimal fraction of
those imported from Mars within Earth’s history.

Fig. 4.2 Number of fragments ejected within the last 4 Ga at temperatures ≤100 °C from Mars or
from the Earth and arriving within 8 Ma on Earth or on Mars (data based on calculations in [26]).
60      G. Horneck et al.

   In experiments simulating impacts comparable to those experienced by the Martian
meteorites, the survival of microbes was tested after subjecting spores of Bacillus
subtilis to accelerations, jerks or shock waves. Accelerations as they occur during
planetary ejections are apparently not a barrier to interplanetary transfer of life. Bacte-
ria are routinely treated with even higher values of acceleration in normal microbio-
logical separation techniques, although with much slower rise times. Ballistic experi-
ments provide rise times equivalent to those estimated for an object receiving escape
velocity during an impact. It has been shown that bacterial spores as well as cells of
Deinococcus radiodurans survived gun shots with accelerations up to 4500 km s-²
(460 000 × g) with a rise time of 10 km (as possibly required for the Martian meteorites studied) the pres-
sure pulse can last up to 20 ms. In shock recovery experiments with an explosive set-
up, the survival of spores of B. subtilis was studied after a shock treatment in the pres-
sure range, which some Martian meteorites have experienced [17]. It was found that
the a substantial fraction of spores (up to 10-4) was able to survive a peak shock pres-
sure of 32 ± 1 GPa and a post-shock temperature of about 250 °C. These data support
the hypothesis that bacterial spores may survive an impact-induced escape process in a
scenario of interplanetary transfer of life. Assuming a mean spore density of 108
spores/g in e.g., desert soil or rock, a 1 kg rock would accommodate approximately
1011 spores, of which up to 107 could survive even extremely high shock pressures
occurring during a meteorite impact. At more moderate shock waves as they would
occur in the spall zone of an impact, even substantial higher survival rates are expected
than in these simulation studies at shock pressures of 32 GPa. To prove this, further
survival studies using such moderate shock waves are required.

4.3 Survival of the Interplanetary Transfer Phase

4.3.1    Space Environment of Interest

Once rocks have been ejected from the surface of their home planet, microbial passen-
gers have to cope with an entirely new set of problems affecting their survival, namely
exposure to the space environment (reviewed in [29-32]). This environment is charac-
terized by a high vacuum, an intense radiation of galactic and solar origin and extreme
temperatures (Table 4.1.).
   In interplanetary space, pressures down to 10-14 Pa prevail. Within the vicinity of a
body, the pressure may significantly increase due to outgassing. In a low Earth orbit,
pressure reaches values of 10-6 to 10-4 Pa. The major constituents of this environment
are molecular oxygen and nitrogen as well as highly reactive oxygen and nitrogen at-
oms.
4 Viable Transfer of Microorganisms in the Solar System and Beyond                   61

Table 4.1. The environment in Earth orbit and of interplanetary space (modified from [34])

Space parameter                   Earth orbit                              Interplanetary space
                                  (≤500km)
Space vacuum
Pressure (Pa)*)                   10-6-10-4                                10-14
Residual gas (part/cm-³)          105 H, 2×106 He, 105 N, 3× 107 O         1H

Solar electromagnetic radiation
Irradiance (W/m²)                 1360                                     Different valuesa)
Spectral range (nm)               Continuum from 2×10-12 to 10² m          Continuum from
                                                                           2×10-12 to 10² m
Cosmic ionizing radiation
Dose (Gy/a) *)                    0.1-3000b)                               ≤ 0.25 c)
Temperature (K)                   100-400a)                                > 4a)
Microgravity (g)                  10-3-10-6                                < 10-6
*) 1 Pa = 10-5 bar, 1 Gy = 100 rad; a) depending on orientation and distance to Sun; b) depending
on altitude and shielding, highest values at high altitudes and in the radiation belts; c) depending
on shielding.

   The radiation environment of our solar system is governed by components of galac-
tic and solar origin. The galactic cosmic radiation entering our solar system is com-
posed of protons (85%), electrons, α-particles (14%) and heavy ions (1%) of charge
Z>2, the so-called HZE particles (high charge Z and high energy E). The solar particle
radiation, emitted in solar wind and during solar particle events, is composed of 90-
95% protons, 5-10% α-particles and a relatively small number of heavier ions. In
interplanetary space, the annual radiation dose amounts to ≤0.25 Gy/a, depending on
mass shielding with the highest dose at 30 g/cm² shielding due to built up secondary
radiation. In the vicinity of the Earth, the radiation dose can increase due to the radia-
tion belts where protons and electrons are trapped by the geomagnetic field.
   The spectrum of solar electromagnetic radiation spans several orders of magnitude,
from short wavelength X-rays to radio frequencies. At the distance of the Earth from
the Sun (1 AU), solar irradiance amounts to 1360 W m-2, the solar constant. Of this
radiation, 45% is attributed to the infrared fraction, 48% to the visible fraction and
only 7% to the ultraviolet range. The extraterrestrial solar spectral UV irradiance has
been measured during several space missions, such as Spacelab 1 and EURECA [33].
   The temperature of a body in space which is determined by the absorption and emis-
sion of energy, depends on its position with respect to the Sun and other orbiting bod-
ies, and also on its surface, size, mass, and albedo (reflectivity). In Earth orbit, the
energy sources include solar radiation (1360 W m-2), the Earth’s albedo (480 W m-2)
and terrestrial radiation (230 W m-2). When orbiting a planet, an object can be shaded
from the Sun as it passes on the planet’s night side. Therefore, the temperature of a
62      G. Horneck et al.

body in space can reach both extremely high and low values. In experiments in Earth
orbit, temperatures between 240 K and 320 K were measured.

4.3.2    Approaches to Studying the Biological Effects of Space

In order to study the survival of resistant microbial forms in the upper atmosphere or in
free space, microbial samples have been exposed in situ by use of balloons, rockets or
space crafts and their responses were investigated after recovery (reviewed in [29, 31]).
For this purpose, several facilities were developed such as the Exposure Device on
Gemini, MEED (microbal ecology exposure device) on Apollo, ES029 on Spacelab 1,
ERA (exobiology radiative Assembly) on EURECA, UV-RAD on Spacelab D2,
BIOPAN on FOTON, and EXPOSE for the International Space Station (ISS) [30, 34]
(see also Fig. 4.6). These investigations were supported by studies in the laboratory, in
which certain parameters of space (high and ultrahigh vacuum, extreme temperature,
UV-radiation of different wavelengths, ionizing radiation) were simulated. The micro-
bial responses (physiological, genetic and biochemical changes) to selected factors
applied separately or in combination were determined.
   Many spore-forming bacteria are found in terrestrial soils and their spores have been
recognized as the hardiest known forms of life on Earth. The developmental pathway
from a vegetatively growing bacterial cell to a spore, i.e. the dormant state, is triggered
by depletion of nutrients in the bacterial cell's environment [35]. In the dormant stage,
spores undergo no detectable metabolism and exhibit a high degree of resistance to
inactivation by various physical insults such as cycles of extreme heat and cold, ex-
treme desiccation including vacuum, UV and ionizing radiation, as well as oxidizing
agents or corrosive chemicals (recently reviewed by Nicholson et al. [32]). The high
resistance of Bacillus endospores is mainly due to two factors: (i) a dehydrated, highly
mineralized core enclosed in a thick protective envelop, the cortex and the spore coat
layers (Fig. 4.3), and (ii) the saturation of their DNA with small, acid-

Fig. 4.3 Electonmicrograph of a spore of B. subtilis with the inner core containing the DNA
surrounded by protective layers, the long axis of the spore is 1.2 µm, the core area 0.25 µm²
(courtesy of S. Pankratz).
4 Viable Transfer of Microorganisms in the Solar System and Beyond                   63

soluble proteins whose binding greatly alters the chemical and enzymatic reactivity of
the DNA [36]. In the presence of appropriate nutrients, spores respond rapidly by ger-
mination and outgrowth, resuming vegetative growth. Hence, spore formation repre-
sents a strategy, by which a bacterium escapes temporally and/or spatially from unfa-
vorable conditions: spores exhibit incredible longevity and can be relocated e.g., by
wind and water, to remote areas. Among the bacterial spores, the endospores of the
genus Bacillus are the best investigated ones [32].
   In addition, a variety of microorganisms exist that are adapted to grow or survive in
extreme conditions of our biosphere. Some of them may be suitable candidates for
studies on microorganisms in space. Examples are endo- or epilithic communities,
consisting of cyanobacteria, algae, fungi and/or lichens, which represent a simple mi-
crobial ecosystem living inside or on rocks [37, 38], or osmophilic microbial assem-
blages that are trapped in evaporite deposits [39, 40], or extremely radiation-resistant
microorganisms, like bacteria of the species D. radiodurans as the most radiation-
resistant bacteria known to exist on Earth today. Although D. radiodurans is non-
sporulating, it can go into a kind of dormancy under certain environmental adverse
conditions such as lack of food, desiccation or low temperatures [41, 42].

4.3.3    Biological Effects of the Vacuum of Space

Because of its extremely dehydrating effect, space vacuum has been considered to be
one of the factors that may prevent interplanetary transfer of life [43]. However, space
experiments have shown that up to 70% of bacterial and fungal spores survive short-
term (e.g., 10 days) exposure to space vacuum, even without any protection [29]. The
chances of survival in space are increased, if the spores are embedded in chemical pro-
tectants such as sugars, or salt crystals, or if they are exposed in thick layers. For exam-
ple, 30% of B. subtilis spores survived nearly 6 years of exposure to space vacuum, if
embedded in salt crystals, whereas approximately 70% survived in the presence of
glucose [21] (Table 4.2.). Sugars and polyalcohols stabilize the structure of

Table 4.2. Survival of spores of B. subtilis after exposure to space vacuum (10-6 to 10-4 Pa)
during different space missions.
  Mission     Duration of Survival fraction at end          Survival fraction at end   References
                vacuum      of exposure in thin               of exposure in thick
               exposure            layers                   layers and presence of
                                                               protective sugars
                                        (%)                           (%)

                             in space          ground       in space      ground
                                               control                    control
SL 1                 10 d 69.3 ± 15.8         85.3 ± 2.6       n.d.         n.d.       [29]
EURECA              327 d 32.1 ± 16.3         32.7 ± 5.6   45.5 ± 0.01   62.7 ± 8.2    [44]
LDEF              2 107 d 1.4 ± 0.8           5.4 ± 2.9    67.2 ± 10.2   77.0 ± 6.0    [21,29]
n.d. = not determined
64      G. Horneck et al.

the cellular macromolecules during vacuum-induced dehydration, leading to increased
rates of survival.
    To determine the protective effects of different meteorite materials, "artificial mete-
orites" were constructed by embedding B. subtilis spores in clay, meteorite dust or
simulated Martian soil [44] and exposing them to the space environment. Crystalline
salt provided sufficient protection for osmophilic microbes in the vegetative state to
survive at least 2 weeks in space [40]. For example, a species of the cyanobacterium
Synechococcus that inhabits gypsum-halite crystals was capable of nitrogen and carbon
fixation and about 5% of a species of the extreme halophile Haloarcula survived after
exposure to the space environment for 2 weeks in connection with a FOTON space
flight.
    The mechanisms of damage due to vacuum exposure are based on the extreme des-
iccation. If not protected by internal or external substances, cells in a vacuum experi-
ence dramatic changes in lipids, carbohydrates, proteins and nucleic acids. Upon desic-
cation the lipid membranes of cells undergo dramatic phase changes from planar bilay-
ers to cylindrical bilayers [45]. The carbohydrates, proteins and nucleic acids undergo
so-called Maillard reactions, i.e. amino-carbonyl-reactions, to give products that be-
come cross-linked eventually leading to irreversible polymerization of the biomolecu-
les [45]. Concomitant with these structural changes are functional changes, including
altered selective membrane permeability, inhibited or altered enzyme activity, de-
creased energy production, alteration of genetic information, etc.
    Vacuum-induced damage to the DNA is especially dramatic, because it may be le-
thal or mutagenic. The mutagenic potential of space vacuum was first demonstrated
during the Spacelab I mission in spores of B. subtilis, which showed an up to tenfold
increased mutation rate over the spontaneous rate [46]. This vacuum-induced muta-
genicity is accompanied by a unique molecular signature of tandem-double base
changes at restricted sites in the DNA [47]. In addition, DNA strand breaks have been
observed to be induced by exposure to space vacuum [48, 49]. Such damage would
accumulate during long-term exposure to space vacuum, because DNA repair is not
active during this state of anhydrobiosis. Survival ultimately depends on the efficiency
of the repair systems after germination.

4.3.4    Biological Effects of Galactic Cosmic Radiation

If not sufficiently shielded by meteorite material, microbes may be affected by the
ionizing components of radiation in space. Especially the heavy primaries of galactic
cosmic radiation, the so-called HZE particles, are the most biologically effective spe-
cies (reviewed in [50, 51]). Because of their low flux (they contribute to approxi-
mately 1% of the flux of particulate radiation in space), methods have been developed
to localize precisely the trajectory of an HZE particle relative to the biological object
and to correlate the physical data of the particle relative to the observed biological
effects along its path. In the Biostack method visual track detectors are sandwiched
between layers of biological objects in a resting state, e.g., B. subtilis spores [52] (Fig.
4.4). This method allows (i) to localize each HZE particle’s trajectory in relation to the
biological specimens; (ii) to investigate the responses of each biological individual hit
separately, in regard to its radiation effects; (iii) to measure the impact parameter b (i.e.
4 Viable Transfer of Microorganisms in the Solar System and Beyond                   65

the distance between the particle track and the sensitive target); (iv) to determine the
physical parameters [charge (Z), energy (E) and linear energy transfer (LET)]; and fi-
nally (v) to correlate the biological effects with each HZE particle parameters.
   The small size of bacterial spores (their cytoplasmic core has a geometrical cross
section of 0.2 to 0.3 µm2) requires to develop special techniques in connection with
the Biostack method. In this case, spores in monolayers are directly mounted on the
track detector. After exposure, the track detector with the spores is etched under micro-
scopical control on the spore-free side only and spores located around a track of an
HZE particle (at a radius of ≤5 µm) are removed by micromanipulation and incubated
each individually in special incubation chambers [53] (Fig. 4.4). Using this micro-
scopical method, the accuracy in determining the impact parameter was ≥0.2 µm, de-
pending on the dip angle of the trajectory. Figure 4.4 shows the frequency of inacti-
vated spores as a function of the impact parameter b. Spores within b ≤0.2 µm were
inactivated by 73%. The frequency of inactivated spores dropped abruptly at b >0.2
µm. However, 15-30% of spores located within 0.2 < b 10 µm) (95% confidence) [54].

Fig. 4.4 Biostack method to localize the effect of single particles (HZE particles) of cosmic
radiation (biological layers are sandwiched between track detectors) and results on the inactiva-
tion probability of spores of B. subtilis as a function of their distance from the particles trajec-
tory, the impact parameter (data from space experiments [54]).
66      G. Horneck et al.

   As shown above, B. subtilis spores can survive even a central hit of an HZE particle
of cosmic radiation. Such HZE particles of cosmic radiation are conjectured to set the
ultimate limit on the survival of spores in space, because they penetrate even thick
shielding. However, since the flux of HZE particles is relatively low, it may last up to
several hundred thousand to one million of years in space until a spore might be hit by
an HZE particle (e.g., iron of LET >100 keV/µm).
   With increasing shielding thickness, e.g., by the outer layers of the meteorite, the
dose rate caused by cosmic radiation goes through a maximum, because the heavy ions
interact with the shielding material and create secondary radiation. Based on experi-
mental data from accelerator experiments with B. subtilis spores [55] and biophysical
models, relating the structure of HZE tracks to the probability of inactivating the
spores [56], an estimated density of the meteorite of 3 g/cm³ (taken from data on Mar-
tian meteorites) and a NASA model on an HZE transport code for cosmic radiation
[57], the dose rates and probabilities for inactivating spores have been calculated be-
hind different shielding thicknesses: the physical dose rates reach a maximum behind a
shielding layer of about 10 cm (30 g/cm² shield thickness); behind 30 cm (90 g/cm²)
the value is approximately the same as obtained without any shielding and only for
higher shielding thicknesses the dose rate reduces significantly (Fig. 4.5) [26]. The
calculations also show that even after 25 Ma in space, a substantial fraction of a spore
population (10-6) would survive the exposure to cosmic radiation if shielded by 2 to
3 m of meteorite material. The calculations are based on the assumption that the rock
may accommodate about 108 spores/g, of which at least 100 spores/g would survive.
The same surviving fraction would be reached after about 600 000 years without

Fig. 4.5 Shielding of spores of B. subtilis against galactic cosmic radiation (GCR) by meteorite
material and survival times (≥ 10-6 survivors) at different depths of the meteorite due to GCR
(dashed line) or to GCR plus natural radioactivity of 0.8 mGy/a (dotted line) (data from [26]).
4 Viable Transfer of Microorganisms in the Solar System and Beyond          67

any shielding, after about 300 000 years behind 10 cm of shielding (maximum dose
rate) and after about 1 Ma behind 1 m of shielding.

4.3.5    Biological Effects of Extraterrestrial Solar UV Radiation

Solar UV radiation has been found to be the most deleterious factor of space, as tested
with dried preparations of viruses, and of bacterial and fungal spores [29]. The full
spectrum of extraterrestrial UV radiation kills unprotected spores of B. subtilis within
seconds [2] and is about thousand times more efficient than the UV radiation at the
Earth’s surface due to the effective protection of our biosphere by the stratospheric
ozone layer [58, 59] (see also Chap. 14, Cockell; Chap. 15, Rettberg and Rothschild).
The reason for this high biological efficiency of extraterrestrial UV is the highly ener-
getic UV-C (190-280 nm) and vacuum UV region (
68      G. Horneck et al.

    Besides cosmic radiation and vacuum, threats to the stability of the DNA of spores
inside meteorites may arise from various sources such as natural radioactivity of the
meteorite, hydrolysis, chemicals, and temperature extremes. The natural radioactivity
of the known Martian meteorites reaches values up to 0.8 mGy/a. It has been calculated
by Mileikowsky et al. [26] that for smaller Martian ejecta (
4 Viable Transfer of Microorganisms in the Solar System and Beyond         69

planet [6, 7, 14]. Therefore, 6 years seem to be by far too short for an interplanetary
transfer of life.

4.4 Survival of the Landing Process
When captured by a planet with an atmosphere, most meteorites are subjected to very
high temperatures during landing. However, because the fall through the atmosphere
takes a few seconds only, the outermost layers form a kind of heat shield and the heat
does not reach the inner parts of the meteorite. During entry, the fate of the meteorite
strongly depends on its size: large meteorites may brake into pieces, however, these
may be still large enough to remain cool inside until hitting the surface of the planet;
medium sized meteorites may obtain a melted crust, whereas the inner part still re-
mains cool; micrometeorites of a few µm in size may tumble through the atmosphere
without being heated at all above 100 °C. Therefore, it is quite possible that a substan-
tial number of microbes can survive the landing process on a planet. However, no
experiments have been done so far to investigate the effects of the landing process
experimentally. Recently, the European Space Agency (ESA) has developed a facility,
called STONE which is attached to the heat shield of a FOTON satellite to test mineral
degradations during landing [64]. This facility might be an ideal tool to study the ef-
fects of landing of bacterial spores embedded in an artificial meteorite.

4.5 Conclusions: On the Likelihood
    of Interplanetary Transfer of Life as a Mode
    of Distribution of Life Throughout the Solar System
Although it is difficult at present to prove definitely that life has been transported
through our solar system, model calculations and experiments at simulation facilities
and in space allow to estimate the chances of resistant microbial forms to survive the
different steps of such a scenario. Experiments in space which were performed on free
platforms since the Apollo era and more sophisticated on the external platform of
Spacelab and on free flying satellites such as LDEF, EURECA, and FOTON, have
given some insight into responses of bacterial spores and other extremophile microor-
ganisms to the parameters of space. The most interesting results are summarized in the
following: (i) extraterrestrial solar UV radiation is thousand times more efficient than
UV at the surface of the Earth and kills within a few seconds 99% of B. subtilis
spores; (ii) space vacuum increases the UV sensitivity of the spores; (iii) although
spores survive extended periods of time in space vacuum (up to 6 years), genetic
changes occur such as increased mutation rates; (iv) after 6 years in space, up to 70%
of bacterial spores survive, if protected against solar UV radiation and dehydration; (v)
spores could escape a hit of a cosmic HZE particle (e.g., iron ion) for up to 1 Ma.
Calculations using radiative transfer models for cosmic rays and biological data from
accelerator experiments have shown that a meteorite layer of 1 m or more effectively
protects bacterial spores against galactic cosmic radiation.
70      G. Horneck et al.

   The data obtained so far on the responses of resistant microorganisms to the envi-
ronment of space support the supposition that space – although it is very hostile to
terrestrial life – is not a barrier for cross-fertilization in the solar system. Ejection of
microbes inside rocks and their transport through the solar system is a feasible process.
If protected against solar UV and galactic cosmic radiation, spores may survive inside
meteorites over extended periods of time.
   However, several aspects justify continued research, e.g., on the survival rate during
the ejection and landing process, on the shelf-life of spores in vacuum, as well as on
the best mechanisms to effectively protect the microorganisms enclosed against the
three steps of interplanetary transfer. Relevant studies to tackle these questions will be
performed in future studies in space and at ground based facilities [34, 65, 66]. For
future research on bacterial spores and other microorganisms in space, ESA is devel-
oping the EXPOSE facility that is to be accomodated outside of the ISS for 1.5 years
(Fig. 4.6). EXPOSE will support long-term in situ studies of microbes in artificial
meteorites, as well as of microbial communities from special ecological niches such as
endolithic and endoevaporitic ecosystems [34]. These experiments on the Responses of
Organisms to the Space Environment (ROSE) include the study of photobiological
processes in simulated radiation climates of planets (e.g., early Earth, early and present
Mars, and the role of the ozone layer in protecting the biosphere from harmful

Fig. 4.6 EXPOSE facility to be mounted on the truss structure of the ISS to study the sensitiv-
ity of organics and microorganisms to space environment.
4 Viable Transfer of Microorganisms in the Solar System and Beyond                     71

Table 4.3. Experiments of the ROSE consortium to study the responses of organisms to space
environment on the EXPOSE facility of the ISS [34]
Code           Objective                                   Assay system

ENDO           Impact of ozone depletion on micro-         1. Endolithic microbial communities
               bial primary producers from sites           2. Mats of cyanobacteria
               under the "ozone hole"                      3. Mats of algae
OSMO           Protection in evaporites against solar      1. Synechococcus in and beneath
               UV and anhydrobiosis                           gypsum-halite
                                                           2. Haloarcula (pigmented and non-
                                                              pigmented) in and beneath NaCl
                                                           3. Haloarcula DNA in KCl
SPORES         Protection of spores by meteorite           1. B. subtilis spores
               material against space (UV, vacuum,         2. Fungal spores
               and ionizing radiation)                     3. Lycopod spores
                                                           All in or beneath meteorite material
PHOTO          Main photoproducts in dry DNA and           1. DNA
               DNA from dry spores                         2. Bacterial spores
PUR            Sensitivity of the biologically effective   1. T7 bacteriophage
               UV radiation to ozone                       2. Phage-DNA
                                                           3. Uracil
SUBTIL         Mutational spectra induced by space         1. B. subtilis spores
               vacuum and solar UV                         2. Plasmid DNA

UV-B radiation), as well as studies of the probabilities and limitations for life to be
distributed beyond its planet of origin (Table 4.3.). Here-to-fore, the results from the
EXPOSE experiments will eventually provide clues to a better understanding of the
processes regulating the interactions of life with its environment.

4.6 Outlook: On the Likelihood of Transport
    of Viable Microorganisms Between Solar Systems
Today more than 50 extrasolar planets have been detected, all of the size of Saturn or
Jupiter (see Chap. 2, Udry and Mayor). The present detection methods – Doppler ef-
fect, astrometry and transit photometry - do not possess sufficient sensitivity to dis-
cover Earth-like planets nor giant planets, orbiting at distances larger than 3 AU from
their sun. For that we have to wait for future planned space missions, e.g., Darwin of
ESA (see Chap. 24, Foing) which are especially designed to detect Earth-like planets by
searching for signatures indicative of a biosphere on a planet, such as atmospheric
oxygen as a suggested biomarker for photosynthetic activity. Nevertheless, the infor-
mation on the existence of extrasolar planets has corroborated the interest in the ques-
tion whether life forms can be transported between the planets or moons of different
solar systems.
72       G. Horneck et al.

   In a follow-on study, Mileikowsky et al. [67] have estimated the likelihood of viable
transfer of life from planets or moons of extrasolar planetary systems to the Earth
where the source planet is located in either the general galactic star field or in a possi-
ble temporary "sibling" cluster born together with the Sun. To leave the planet, an
impact ejection mechanism has been assumed similar as described in the studies of
viable transfer within our solar system [26]. Because the information on the existence
and frequency of Earth-like planets is still missing – and probably for many years to
come -, they have based their calculations on specific estimations of the unknown
circumstances as described below.
   Major parameters of the calculations are (i) the star density and (ii) the relative
speed between the emitting and receiving planetary system. In the case of the general
galactic field, the present numerical density of stars in the solar neighborhood was
used, namely 1 star per 10 pc³ = 0.1 pc-3 (1pc = parsec = 3.26 light years). This is based
on the generally held opinion of astronomers that the general star field during the youth
of our solar system was probably quite similar to today's. The distribution of speeds of
the stars in our galaxy is a very good Maxwellian approximation with about 20 km/s as
the most common speed and 0.5 km/s being orders of magnitude less frequent. In the
case of a possible cluster of "sibling" stars born together with our Sun, the cluster
would last only limited time after its formation from the molecular cloud, before the
stars disperse in all directions into the general galactic star field. However, the dis-
tances between the stars within the cluster would be much smaller than in the general
galactic star field. Furthermore, the relative speed between the members of a cluster is
less than 1 km/s which is much lower than the most frequent one between stars in the
general galactic star field. To calculate the number n of hits on Earth by ejecta from
extrasolar systems, the following formula was used [67]:

                                         ν0                                                 (4.1)
                                            2

     n ≈ 2 ⋅ 10   −15
                        T Rt
                             N   1 − e − 2σ 2   
                             σ                  
                                                

      With                       N                      =                      Nd×fps×fTP×fHZM×fi,

      Nd = numerical density of stars in pc-3; fps = fraction of stars with planetary systems; fTP
      = fraction of planetary systems with terrestrial-like planets; fHZM = fraction of terrestrial-
      like planets orbiting in zones habitable for microbes; fi = fraction of other factors; T =
      time period studied comprising escape from a planet, expulsion into interstellar space,
      capture by our solar system, orbiting in our solar system, capture by the Earth, and hitting
      the Earth (Ma); t = time period for microbial survival inside ejecta in space (Ma); R = rate
      of expulsion from extrasolar systems into interstellar space (Ma-1); σ = dispersion in the
      Maxwellian distribution of the velocities of the stars (km/s); ν 0= 0.5 km/s.
   For establishing the formula, it is not necessary to know either anything about the
complete planetary configuration of the extrasolar systems or about the terrestrial-like
planets. But, for each open entity, selected values have to be given. Nd = 0.1 pc-3 and σ
= 20 pc/Ma are the values known for our galaxy. For a possible temporary "sibling"
cluster around our Sun, Nd and σ are of course unknown; therefore, the estimations
are based on observations of relatively new-born clusters such as Hyades. t is deter-
4 Viable Transfer of Microorganisms in the Solar System and Beyond           73

mined by biology as discussed in 4.3.4 and in [26]. However, several unknowns render
the calculations difficult, such as (i) the fraction of stars in our galaxy with planetary
systems; (ii) the fraction of those with habitable zones (see Chap. 3, Franck et al.); (iii)
the fraction of those with Earth-like planets; and (iv) the fraction of those that have
developed life (e.g., microorganisms, based on DNA/RNA/proteins with high resis-
tance to the hostile environment of space). To overcome this problem, Mileikowsky et
al. [67] have undertaken the following steps: (i) they have chosen for all fractions
mentioned above the most favorable and overoptimistic value for viable transfer,
namely 1; (ii) they have chosen a heavy bombardment in the early stage of the extraso-
lar system being orders of magnitude more intense than that in the early stage of our
solar system; (iii) they have chosen very long survival times (e.g., 1-100 Ma) of mi-
crobes in ejecta in space. Table 4.4. shows that the transfer of ejecta from an extrasolar
system to the Earth calculated by use of Eq. (4.1), is strongly enhanced by high densi-
ties, i.e. short distances (e.g., less than 1 pc) and/or by a very low relative speed be-
tween the emitting planetary system and the receiving system (e.g., less than 0.5 km/s)
[67, 68].
   In case of the conditions of the general galactic star field, n, the number of ejecta
from all planetary systems of the galaxy, reaching the Earth within the early 500 Ma is
utterly small, namely about 10-9. Even if the bombardment by comets and asteroids
would be thousand times more intense (given by term R), no ejecta would reach the
Earth either, with n being about 10-5. In the case of clusters of "sibling" stars, assuming
the conditions as observed in the Hyades cluster, which has at its present age of 625
Ma N = 2 and σ = 0.25 km/s, n would be about 10-4. Even assuming a 20 times higher
star density than observed in Hyades, the number of extrasolar ejecta hitting the Earth
is still far below 1. Hence, assuming impact ejecta as the mode of transportation of
putative life from an extrasolar planet to the Earth, the data show that from the general
galactic star field no life-bearing ejecta have reached the Earth within the first 05-0.6
Ga. If the sun was part of a group of stars in a "sibling" cluster, viable transfer from
one of the sister systems cannot be completely ruled out, however, the probability
remains very low.
   The longest survival time assumed is t=100 Ma (Table 4.4.). Recently, revival of
bacterial spores from a 250 Ma old salt deposit has been claimed [20]. However, as
discussed above, this finding is still controversial: it has been argued that the spores
may be the result of recent contamination, which have penetrated the salt through small
invisible cracks. This objection is mainly based on the high genetic similarity of the
revived spores with contemporary species of Bacillus maremortis, which live in the
very saline Dead Sea. However, it can be deduced from Eq. (4.1) that the number of
viable transfers from planets of the general galactic field to the Earth would remain
low even on the base of a survival time in space of 250 Ma.
   From this example calculated for the case of extrasolar ejecta reaching the Earth, it
can be concluded that the probabilities are too low to allow transport of viable micro-
organisms from one solar system to another by impact ejecta. The chances for such
interstellar exchange of life may be increased for very close sister solar systems which
are born from one parent molecular cloud, where they form a cluster of "sibling" stars
all with planetary bodies, which all – in these calculations – were overoptimistically
assumed to be populated by microbes.
74          G. Horneck et al.

Table 4.4. Number of impacts n on Earth during time interval T assuming maximum allowed
microbial travelling time t from ejection to impact, expulsion rate R, density if expelling extraso-
lar systems N, and stellar velocity dispersion σ from the general galactic star field and from
clusters of "sibling"stars

Case                                T        t         R         N         σ         n
                                    (Ma)     (Ma)      (Ma-1)    (pc-3)    (km/s)
From all planetary systems in the 500        10        108       0.1       20        1.6×10-9
general galactic star field
                                  500        100       1011      0.1       20        1.6×10-5
Our Sun within a cluster of         100      1         108       40        0.2       3.8×10-3
"sibling" stars
                                    100      10        108       40        0.2       3.8×10-2
Our Sun within cluster like         625      1         108       2         0.2       8.6×10-4
Hyades

4.7 References
       1     J. Oro, Nature 190, 389 (1961).
       2     G. Horneck, A. Brack, in: S.L. Bonting (Ed.) Advances in Space Biology and Medi-
             cine, Vol. 2, JAI Press, Greenwich, CT, 1992, pp. 229.
       3     C.F. Chyba, T.C. Owen, W.-H. Ip, in: T. Gehrels (Ed.) Hazards due to Comets and
             Asteroids, Univ. of Arizona Press, Tucson, 1994, pp. 9.
       4     H.J. Melosh, A.M. Vickery, Nature 338, 487 (1989).
       5     V.R. Oberbeck, G. Fogleman, Origins Life Evol. Biosphere 20, 181 (1990).
       6     H.J. Melosh, Nature 332, 688 (1988).
       7     B. Gladman, J.A. Burns, M. Duncan, P. Lee, H. Levinson, Science 271, 1387 (1996).
       8     P. Warren, Icarus 111, 338 (1994).
       9     O. Eugster, A. Weigel, E. Polnau, Geochim. Cosmochim. Acta 61, 2749 (1997).
       10    H. Richter, Schmidts Jahrbuch Ges. Med. 126, 243 (1865).
       11    S. Arrhenius, Die Umschau 7, 481 (1903).
       12    G. Horneck, Planet. Space Sci. 43, 189 (1995).
       13    D.S. McKay, E.K. Gibson, K.L. Thomas-Keprta,                H. Vali, C.S. Romanek,
             S.J. Clemett, X.D.F. Chillier, C.R. Maedling, R.N. Zare, Science 273, 924 (1996).
       14    M.A. Moreno, Nature 336, 209 (1988).
       15    J.D. O'Keefe, T.J. Ahrens, Science 234, 346 (1986).
       16    A.M. Vickery, H.J. Melosh, Science 237, 738 (1987).
       17    G. Horneck, D. Stöffler, U. Eschweiler, U. Hornemann, Icarus 149, 285 (2001).
       18    P. Weber, J.M. Greenberg, Nature 316, 403 (1985).
       19    R.J. Cano, M. Borucki, Science 268, 1060 (1995).
       20    R.H. Vreeland, W.D. Rosenzweig, D.W. Powers, Nature 407, 897(2000).
       21    G. Horneck, H. Bücker, G. Reitz, Adv. Space Res. 14(10), 41 (1994).
       22    M. Schidlowski, Nature 333, 313 (1988).
       23    C.R. Woese, O. Kandler, M.L. Wheelis, Proc. Natl. Acad. Sci. USA 87, 4576 (1990).
4 Viable Transfer of Microorganisms in the Solar System and Beyond               75

24   D. Stöffler, 31st Lunar and Planetary Science Conference, Houston, Texas, CD ROM
     #1170, 2000.
25   H.J. Melosh, Icarus 59, 234 (1984).
26   C. Mileikowsky,      F.A. Cucinotta,      J.W. Wilson,      B. Gladman,    G. Horneck,
     L. Lindegren, H.J. Melosh, H. Rickman, M. Valtonen, J.Q. Zheng, Icarus 145, 391
     (2000).
27   C. Mileikowsky, E. Larsson, B. Eiderfors, 25th General Assembly of the European
     Geophysical Society, The Hague, 19-23 April 1999.
28   R.M.F. Mastrapa, H. Glanzberg, J.N. Head, H.J. Melosh, W.L. Nicholson, 31st Lunar
     and Planetary Science Conference, Houston TX, CD ROM #2045, 2000.
29   G. Horneck, Origins Life Evolut. Biosphere 23, 37 (1993).
30   G. Horneck, Adv. Space Res. 22(3), 317 (1998).
31   G. Horneck, in: L.J. Rothschild, A. Lister (Eds.) Evolution on Planet Earth: The Im-
     pact of the Physical Environment, Academic Press, in press.
32   W.L. Nicholson, N. Munakata, G. Horneck, H.J. Melosh, P. Setlow, Microbiol. Mo-
     lec. Biol. Rev. 64, 548 (2000).
33   D. Labs, H. Neckel, P.C. Simon, G. Thuillier, Solar Phys. 107, 203 (1987).
34   G. Horneck, D.D. Wynn-Williams, R. Mancinelli, J. Cadet, N. Munakata, G. Rontó,
     H.G.M. Edwards, B. Hock, H. Wänke, G. Reitz, T. Dachev, D.P. Häder, C. Brillouet,
     Proc. 2nd Europ. Symp. on Utilisation of the Internat. Space Station, ESTEC, Noord-
     wijk, ESA SP-433, 1999, pp. 459.
35   P.J. Piggot, C.P. Jr. Moran, P. Youngman, (Eds.) Regulation of Bacterial Differen-
     tiation. American Society for Microbiology, Washington, D.C., 1994.
36   P. Setlow, Ann. Rev. Microbiol. 49, 29 (1995).
37   E.I. Friedmann, Science 215, 1045 (1982).
38   D.D. Wynn-Williams, Microbial Ecology 31, 177 (1996).
39   L.J. Rothschild, L.J. Giver, M.R. White, R.L. Mancinelli, J. Phycol. 30, 431 (1994).
40   R.L. Mancinelli, M.R. White, L.J. Rothschild, Adv. Space Res. 22(3), 327 (1998).
41   B.E.B. Moseley, in: K.C. Smith (Ed.) Photochemical and Photobiological Reviews
     Vol. 7, Plenum, New York, 1983, pp. 223.
42   V. Mattimore, J.R. Battista, J. Bacteriol. 178, 633 (1996).
43   M.D. Nussinov, S.V. Lysenko, Origins Life Evol. Biosphere 13, 153 (1983).
44   G. Horneck, U. Eschweiler, G. Reitz, J. Wehner, R. Willimek, K. Strauch, Adv. Space
     Res. 16(8), 105 (1995).
45   C.S. Cox, Origins Life Evol. Biosphere 23, 29 (1993).
46   G. Horneck, H. Bücker, G. Reitz, H. Requardt, K. Dose, K.D. Martens,
     H.D. Mennigmann, P. Weber, Science 225, 226 (1984).
47   N. Munakata, M. Saitou, N. Takahashi, K. Hieda, F. Morohoshi, Mutat. Res. 390,
     189 (1997).
48   K. Dose, A. Bieger-Dose, M. Labusch, M. Gill, Adv. Space Res. 12(4), 221 (1992).
49   K. Dose, A. Bieger-Dose, R. Dillmann, M. Gill, O. Kerz, A. Klein, H. Meinert,
     T. Nawroth, S. Risi, C. Stridde, Adv. Space Res. 16(8), 119 (1995).
50   G. Horneck, Nucl. Tracks Radiat. Meas. 20, 185 (1992).
51   J. Kiefer, M. Kost, K. Schenk-Meuser, in: D. Moore, P. Bie, H. Oser (Eds.) Biologi-
     cal and Medical Research in Space, Springer, Berlin, 1996, pp. 300.
52   H. Bücker, G. Horneck, in: O.F. Nygaard, H.I. Adler, W.K. Sinclair (Eds.) Radiation
     Research, Academic Press, New York, 1975, pp. 1138.
53   M. Schäfer, H. Bücker, R. Facius, D. Hildebrand, Rad. Effects 34, 129 (1977).
54   R. Facius, G. Reitz, M. Schäfer, Adv. Space Res. 14(10), 1027 (1994).
76        G. Horneck et al.

     55    K. Baltschukat, G. Horneck, Radiat. Environm. Biophys. 30, 87 (1991).
     56    F.A. Cucinotta, J.W. Wilson, R. Katz, W. Atwell, G.D. Badhwar, M.R. Shavers, Adv.
           Space Res. 18(12), 183 (1995).
     57    J.W. Wilson, L.W. Townsend, J.L. Shinn, F.A. Cucinotta, R.C. Costen, F.F. Badavi,
           S.L. Lamkin, Adv. Space Res. 10(10), 841 (1994).
     58    G. Horneck, P. Rettberg, E. Rabbow, W. Strauch, G. Seckmeyer, R. Facius,
           G. Reitz, K. Strauch, J.U. Schott, J. Photochem. Photobiol. B: Biol. 32, 189 (1996).
     59    P. Rettberg, G. Horneck, W. Strauch, R. Facius, G. Seckmeyer, Adv. Space Res.
           22(3), 335 (1998).
     60    C. Lindberg, G. Horneck, J. Photochem. Photobiol. B:Biol. 11, 69 (1991).
     61    B. Gladman, J. A. Burns, Science 274, 161 (1996).
     62    G. Horneck, Planet. Space Sci. 48, 1053 (2000).
     63    T. Lindahl, Nature 362, 709 (1993).
     64    A. Brack, ESA Bulletin 101, 101 (2000).
     65    G. Horneck, in: P. Ehrenfreund, C. Krafft, H. Kochan, V. Pirronello (Eds.) Labora-
           tory Astrophysics and Space Research, Kluwer, Dordrecht, 1999, pp. 667.
     66    G. Horneck, G. Reitz, P. Rettberg, M. Schuber, H. Kochan, D. Möhlmann,
           L. Richter, H. Seidlitz, Planet. Space Sci. 48, 507 (2000).
     67    C. Mileikowsky,      F.A. Cucinotta,     J.W. Wilson,       B. Gladman,   G. Horneck,
           L. Lindegren, H.J. Melosh, H. Rickman, M. Valtonen, J.Q. Zheng, in preparation.
     68    M.J. Valtonen, K.A. Innanen, Astrophys. J. 255, 367 (1982).
You can also read